arXiv:1704.00281v1 [math.LO] 2 Apr 2017 adAayi aebe aebfr,eg sfollows: as e.g. before, made been have Analysis dard eut ntecmuainlcneto osadr nlssas Analysis in Nonstandard results Wattenberg’s of study content to computational is the paper on this of results aim the nutshell, a In aigtefloigipratobservation. important following car the separated making (and below treated be will Bishop’s cases from both notion clear, foundational become the as or h nrdcino atnegsppr[7 nldstefloigst following the includes [37] paper Wattenberg’s of introduction The content. iemne ttmnsmyb on n[,1,2–5 2 4 6 3 36, 34, 32, 21–25, 11, [9, word in the found interpret may be may statements Like-minded fMteais hn University Ghent Mathematics, of is fall, of First oee,Wtebr osfrhrta oto h aforement the of most than further goes Wattenberg However, uihCne o ahmtclPiooh,LUMnc,Ger Munich, LMU Philosophy, Mathematical for Center Munich -aladdress E-mail e od n phrases. and words Key OSADR NLSSADCONSTRUCTIVISM! AND ANALYSIS NONSTANDARD h lrpwrcntuto opoueamdl[ nteothe the On ] . upon non- . [. does highly model nonstandard it a is hand, produce as analysis to depending construction nonstandard ultrapower suspect, that the somewhat held thus been constructive, often has It ups fti ae st netgt hs da yexamining by ideas these investigate 303]) to p. ([37, is The examples. paper been ] several . has this . [. of author of constructivism. areas the and purpose unlikely analysis time rather nonstandard two some - between For mathematics affinity apparent an paper. by speculative struck a is This xeddnme e ecnpoedwt xlctcalculations. explicit with 31]) p. proceed [1, can original: in we (Emphasis number extended title mlct lisrgrigtecntutv ttso h a Wattenberg’ the of of status some Part constructive of Standard the incorrectness regarding claims the implicit) Analysi establish Nonstandard we involving hand, modifica y (not slight to theorems seen only are constructive with Analysis Nonstandard resear hand, in recent one theorems of Wattenberg’s On light in connection. work mentioned Wattenberg’s con study and We Analysis Nonstandard matics. between connection possible a Abstract. osadr nlssadConstructivism? and Analysis Nonstandard similar : [email protected] lottodcdsao atnegpbihdapprwt th with paper a published Wattenberg ago, decades two Almost fNntnadAnalysis. Nonstandard of bevtoscnenn h osrciecneto Nonstan- of content constructive the concerning observations constructive osadr nlss osrciemteais computa mathematics, constructive Analysis, Nonstandard praxis . 1. A SANDERS SAM Introduction srmral osrcie aigthe having constructive; remarkably is stemisra/lsia oin‘effective’, notion mainstream/classical the as osrcieAnalysis Constructive nwihh pcltson speculates he which in xioms ) nteother the On s). edeetv and effective ield tutv mathe- structive ho h afore- the on ch epii and (explicit s in oeof some tion, Transfer ay&Department & many oe uhr by authors ioned n[28–31]. in efully). ] h reader The 8]. ih frecent of light [].A will As ([5]). atement: and e r tional 2 AND CONSTRUCTIVISM!

Despite an essential nonconstructive kernel, many nonstandard ar- guments are constructive until the final step, a step that frequently involves the standard part map. ([37, p. 303]) This observation is similar to Osswald’s local constructivity. In particular, Osswald has qualified the observation from the above quotes as Nonstandard Analysis is locally constructive, to be understood as the fact that the mathematics performed in the nonstandard world is highly constructive1. By contrast, the nonstandard axioms (Transfer and Standard Part) needed to ‘jump between’ the nonstandard world and usual mathematics, are highly non-constructive in general. Osswald discusses local constructivity in [38, 7], [21, 1-2], or [22, 17.5]. § § § The results in [28–31] vindicate the Wattenberg and Osswald view in that com- putational content is extracted from theorems of ‘pure’ Nonstandard Analysis, i.e. formulated solely with the nonstandard definitions (of continuity, Riemann integra- tion, compactness, et cetera) rather than the ‘ε-δ’ definitions. With this choice, one only works in the nonstandard universe, avoiding the non-constructive step from and to the standard/usual universe (requiring Transfer and Standard Part). In this paper, we show that Wattenberg’s results from [37] yield effective and constructive results with only slight modification. However, we also establish the incorrectness of Wattenberg’s claims regarding the constructive status of the non- standard axioms Transfer and Standard Part. In contrast to Wattenberg, we shall work in Nelson’s axiomatic approach to Nonstandard Analysis (See Section 2), but this change of framework will have no real impact on our results or Wattenberg’s.

2. Internal and its fragments In this section, we discuss Nelson’s theory, first introduced in [18], and its fragments P and H from [3]. The latter fragments are essential to our enterprise, especially Theorem 2.4 below. 2.1. Internal set theory 101. In Nelson’s syntactic approach to Nonstandard Analysis ([18]), as opposed to Robinson’s semantic one ([26]), a new predicate ‘st(x)’, read as ‘x is standard’ is added to the language of ZFC, the usual foundation of mathematics. The notations ( stx) and ( sty) are short for ( x)(st(x) ... ) and ( y)(st(y) ... ). A formula∀ is called internal∃ if it does not∀ involve ‘st’,→ and external∃ otherwise.∧ The three external axioms Idealisation, Standard Part, and Transfer govern the new predicate ‘st’. These axioms are respectively defined2 as: (I) ( st finx)( y)( z x)ϕ(z,y) ( y)( stx)ϕ(x, y), for internal ϕ. (S) (∀stx)( st∃y)( ∀stz)∈(z x ϕ→(z))∃ z∀ y , for any ϕ. ∀ ∃ ∀ ∈ ∧ ↔ ∈ (T) ( stt) ( stx)ϕ(x, t) ( x)ϕ(x, t) , where ϕ(x, t) is internal, and only has  free∀ variables∀ t, x. → ∀   The system IST is (the internal system) ZFC extended with the aforementioned external axioms; The former is a conservative extension of ZFC for the internal language, as proved in [18]. Clearly, the extension from ZFC to IST can be done for other systems, and we are interested in the formalisations of (classical and

1The mathematics performed in the nonstandard world usually amounts to merely manipulat- ing sums and products of nonstandard length. 2The superscript ‘fin’ in (I) means that x is finite, i.e. its number of elements are bounded by a natural number. NONSTANDARD ANALYSIS AND CONSTRUCTIVISM! 3 intuitionistic) arithmetic, namely Peano and Heyting arithmetic. In this regard, the systems H and P from [3], also sketched in the next sections, are nonstandard extensions of the (internal) logical systems E-HAω and E-PAω, respectively Heyting and Peano arithmetic in all finite types and the axiom of extensionality. We refer to [12, 3.3] for the exact definitions of the (mainstream in mathematical ) systems§E-HAω and E-PAω.

2.2. The classical system P. In this section, we introduce the system P, a con- servative extension of Peano arithmetic with fragments of Nelson’s IST. ω ω To this end, recall that E-PA ∗ and E-HA ∗ are the definitional extensions of E-PAω and E-HAω with types for finite sequences, as in [3, 2]. For the former systems, we require some notation. § ω ω Notation 2.1 (Finite sequences). The systems E-PA ∗ and E-HA ∗ have a dedi- cated type for ‘finite sequences of objects of type ρ’, namely ρ∗. Since the usual coding of pairs of numbers goes through in both, we shall not always distinguish ρ ρ between 0 and 0∗. Similarly, we do not always distinguish between ‘s ’ and ‘ s ’, where the former is ‘the object s of type ρ’, and the latter is ‘the sequence ofh typei ρ ρ∗ with only element s ’. The empty sequence for the type ρ∗ is denoted by ‘ ρ’, usually with the typing omitted. Furthermore, we denote by ‘ s = n’ the lengthhi of ∗ ρ ρ ρ ρ | | the finite sequence s = s0,s1,...,sn 1 , where = 0, i.e. the empty sequence h ∗ ∗ − i |hi| has length zero. For sequences sρ ,tρ , we denote by ‘s t’ the concatenation of s and t, i.e. (s t)(i)= s(i) for i< s and (s t)(j)= t( s ∗ j) for s j < s + t . ∗ ∗ | | ∗ | |− | |≤ | | | | For a sequence sρ , we define sN := s(0),s(1),...,s(N) for N 0 < s . For a 0 ρ h i 0| | sequence α → , we also write αN = α(0), α(1),...,α(N) for any N . By way ∗ ρ ρ h i of shorthand, q Q abbreviates ( i < Q )(Q(i) =ρ q). Finally, we shall use ∈ σ0 σ∃k | | x,y,t,... as short for tuples x0 ,...xk of possibly different type σi. ω We can now introduce E-PAst∗. We use the same definition as [3, Def. 6.1], where ω ω E-PA ∗ is the definitional extension of E-PA with types for finite sequences from ω [3, 2]. The set ∗ is the collection of all the terms in the language of E-PA ∗. § T ω ω st Definition 2.2. The system E-PAst∗ is defined as E-PA ∗ + st∗ + IA , where st∗ consists of the following basic axiom schemas. T T (1) The schema3 st(x) x = y st(y), ∧ → 4 (2) The schema providing for each closed term t ∗ the axiom st(t). (3) The schema st(f) st(x) st(f(x)). ∈ T ∧ → The external induction axiom IAst is as follows. Φ(0) ( stn0)(Φ(n) Φ(n + 1)) ( stn0)Φ(n). (IAst) ∧ ∀ → → ∀ Secondly, we introduce some essential fragments of IST studied in [3]. Definition 2.3. [External axioms of P]

3 ω∗ The language of E-PAst contains a symbol stσ for each finite type σ, but the subscript is ∗ essentially always omitted. Hence Tst is an axiom schema and not an axiom. 4A term is called closed in [3] (and in this paper) if all variables are bound via lambda ab- straction. Thus, if x,y are the only variables occurring in the term t, the term (λx)(λy)t(x,y) is closed while (λx)t(x,y) is not. The second axiom in Definition 2.2 thus expresses that stτ (λx)(λy)t(x,y) if (λx)(λy)t(x,y) is of type τ. We usually omit lambda abstraction for brevity. 4 NONSTANDARD ANALYSIS AND CONSTRUCTIVISM!

(1) HACint: For any internal formula ϕ, we have ∗ st ρ st τ st ρ τ st ρ τ ( x )( y )ϕ(x, y) F → ( x )( y F (x))ϕ(x, y), (2.1) ∀ ∃ → ∃ ∀ ∃ ∈ (2) I: For any internal formula ϕ, we have ∗ ( stxσ )( yτ )( zσ x)ϕ(z,y) ( yτ )( stxσ)ϕ(x, y), ∀ ∃ ∀ ∈ → ∃ ∀ ω (3) The system P is E-PAst∗ + I + HACint.

Note that I and HACint are fragments of Nelson’s axioms Idealisation and Stan- dard part. By definition, F in (2.1) only provides a finite sequence of witnesses to ( sty), explaining its name Herbrandized . ∃ The system P is connected to E-PAω by Theorem 2.4. The latter (which is not present in [3]) expresses that we may obtain effective results as in (2.3) from any theorem of Nonstandard Analysis which has the same form as in (2.2).

Theorem 2.4. If ∆int is a collection of internal formulas and ψ is internal, and P + ∆ ( stx)( sty)ψ(x,y,a), (2.2) int ⊢ ∀ ∃ 4 then one can extract from the proof a sequence of closed terms t in ∗ such that T ω E-PA ∗ + ∆ ( x)( y t(x))ψ(x,y,a). (2.3) int ⊢ ∀ ∃ ∈ Proof. See e.g. [30, 2] or [29, 2].  § § For the rest of this paper, the notion ‘normal form’ shall refer to a formula as in (2.2), i.e. of the form ( stx)( sty)ϕ(x, y) for ϕ internal. It is shown in [29–31] that the scope of Theorem∀ 2.4 includes∃ the ‘Big Five’ systems of Reverse Mathematics and the associated ‘zoo’ from [7]. Finally, the previous theorems do not really depend on the presence of full Peano arithmetic. We shall study the following subsystems. Definition 2.5. [Weaker Systems] ω ω (1) Let E-PRA be the system defined in [13, 2] and let E-PRA ∗ be its defi- nitional extension with types for finite sequences§ as in [3, 2]. (2) (QF-ACρ,τ ) For every quantifier-free internal formula ϕ(x,§ y), we have ρ τ ρ τ ρ ( x )( y )ϕ(x, y) ( F → )( x )ϕ(x, F (x)) (2.4) ∀ ∃ → ∃ ∀ ω ω 1,0 (3) The system RCA0 is E-PRA + QF-AC . ω The system RCA0 is the ‘base theory of higher-order Reverse Mathematics’ as introduced in [13, 2]. We permit ourselves a slight abuse of notation by also § ω 1,0 ω referring to the system E-PRA ∗ + QF-AC as RCA0 . ω Corollary 2.6. The previous theorem and corollary go through for P and E-PA ∗ ω 1,0 ω replaced by P E-PRA ∗ + ∗ + HAC + I + QF-AC and RCA . 0 ≡ Tst int 0 ω Proof. The proof of [3, Theorem 7.7] goes through for any fragment of E-PA ∗ which includes EFA, sometimes also called I∆0 +EXP. In particular, the exponential function is (all what is) required to ‘easily’ manipulate finite sequences.  We now discuss the Standard Part principle Ω-CA, a very practical consequence of the axiom HACint. Intuitively speaking, Ω-CA expresses that we can obtain the standard part (in casu G) of Ω-invariant nonstandard objects (in casu F ( ,M)). Note that we write ‘N Ω’ as short for st(N 0). · ∈ ¬ NONSTANDARD ANALYSIS AND CONSTRUCTIVISM! 5

(σ 0) 0 0 Definition 2.7. [Ω-invariance] Let F × → be standard and fix M Ω. Then F ( ,M) is Ω-invariant if ∈ · ( stxσ)( N 0 Ω) F (x, M)= F (x, N) . (2.5) ∀ ∀ ∈ 0 (σ 0) 0 0 Principle 2.8 (Ω-CA). Let F × →  be standard and fix M Ω. For every σ 0 ∈ Ω-invariant F ( ,M), there is a standard G → such that · ( stxσ)( N 0 Ω) G(x)= F (x, N) . (2.6) ∀ ∀ ∈ 0 The axiom Ω-CA provides the standard part of a nonstandard object, if the latter is independent of the choice of nonstandard number used in its definition. The following theorem is not new, but highly instructive in light of Remark 2.10.

Theorem 2.9. The system P0 proves Ω-CA. Proof. Let F ( ,M 0) be Ω-invariant, i.e. we have · ( stxσ)( N 0,M 0 Ω) F (x, M)= F (x, N) . (2.7) ∀ ∀ ∈ 0 By underspill (See Theorem 2.13 below), we immediately obtain that st σ st 0 0 0 ( x )( k )( N ,M k) F (x, M)=0 F (x, N) . ∀ ∃ ∀ ≥ ∗ σ 0 Now apply HACint to obtain standard Φ →  such that  ( stxσ)( k0 Φ(x))( N 0,M 0 k) F (x, M)= F (x, N) . ∀ ∃ ∈ ∀ ≥ 0 Next, define Ψ(x) := maxi< Φ(x) Φ(x)(i) and note that  | | ( stxσ)( N 0,M 0 Ψ(x)) F (x, M)= F (x, N) . ∀ ∀ ≥ 0 Finally, put G(x) := F (x, Ψ(x)) and note that Ω-CA follows.  

Finally, we consider the following remark on how HACint and I are used.

Remark 2.10 (Using HACint and I). By definition, HACint produces a functional ∗ σ τ F → which outputs a finite sequence of witnesses. However, HACint provides an actual witnessing functional assuming (i) τ = 0 in HACint and (ii) the for- st σ 0 0 mula ϕ from HACint is ‘sufficiently monotone’ as in: ( x ,n ,m ) [n 0 m ∀ σ 0 ≤ σ ∧ ϕ(x, n)] ϕ(x, m) . Indeed, in this case one simply defines G → by G(x ) := → st σ maxi< F (x) F (x)(i) which satisfies ( x )ϕ(x, G(x)), as was done in the proof of | |  ∀ Theorem 2.9. To save space in proofs, we will sometimes skip the (obvious) step involving the maximum of finite sequences, when applying HACint. We assume the same convention for terms obtained from Theorem 2.4, and applications of the contraposition of Idealisation I. 2.3. The constructive system H. In this section, we define the system H, the constructive counterpart of P. The system H was first introduced in [3, 5.2], and constitutes a conservative extension of Heyting arithmetic E-HAω by [3, Cor.§ 5.6]. ω ω st ω Similar to Definition 2.2, we define E-HAst∗ as E-HA ∗ + st∗ +IA , where E-HA ∗ ω T is essentially just E-PA ∗ without the law of excluded middle. Furthermore, define ω st H E-HAst∗ + HAC + I + NCR + HIP st + HGMP , ≡ ∀ where HAC is HACint without any restriction on the formula, and where the remain- ing axioms are defined in the following definition. Definition 2.11. [Three axioms of H] 6 NONSTANDARD ANALYSIS AND CONSTRUCTIVISM!

(1) HIP st ∀ st st st st [( x)φ(x) ( y)Ψ(y)] ( y′)[( x)φ(x) ( y y′)Ψ(y)], ∀ → ∃ → ∃ ∀ → ∃ ∈ ω where Ψ(y) is any formula and φ(x) is an internal formula of E-HA ∗. (2) HGMPst st st [( x)φ(x) ψ] ( x′)[( x x′)φ(x) ψ] ∀ → → ∃ ∀ ∈ → ω where φ(x) and ψ are internal formulas in the language of E-HA ∗. (3) NCR ∗ τ st ρ st ρ τ ( y )( x )Φ(x, y) ( x )( y )( x′ x)Φ(x′,y), ∀ ∃ → ∃ ∀ ∃ ∈ ω where Φ is any formula of E-HA ∗ Intuitively speaking, the first two axioms of Definition 2.11 allow us to perform a number of non-constructive operations (namely Markov’s principle and indepen- dence of premises) on the standard objects of the system H, provided we introduce a ‘Herbrandisation’ as in the consequent of HAC, i.e. a finite list of possible wit- nesses rather than one single witness. Furthermore, while H includes Idealisation I, one often uses the latter’s classical contraposition, explaining why NCR is useful (and even essential) in the context of intuitionistic logic. Surprisingly, the axioms from Definition 2.11 are exactly what is needed to con- vert nonstandard definitions (of continuity, integrability, convergence, et cetera) into the normal form ( stx)( sty)ϕ(x, y) for internal ϕ, as done in [29, 3.1]. The latter normal form plays∀ an∃ equally important role in the constructive§ case as in the classical case by the following theorem.

Theorem 2.12. If ∆int is a collection of internal formulas, ϕ is internal, and H + ∆ stx sty ϕ(x,y,a), (2.8) int ⊢ ∀ ∃ then one can extract from the proof a sequence of closed terms t in ∗ such that T ω E-HA ∗ + ∆ x y t(x) ϕ(x,y,a). (2.9) int ⊢ ∀ ∃ ∈ Proof. Immediate by [3, Theorem 5.9]. Note that in the latter, just like in the proof of Corollary 2.4, stx sty ϕ(x,y,a) is proved to be ‘invariant’ under a suitable ∀ ∃ syntactic translation.  Finally, we point out some very useful principles to which we have access.

Theorem 2.13. The systems P, P0, and H prove and underspill, i.e. ( stxρ)ϕ(x) ( yρ) st(y) ϕ(y) and ( xρ) st(x) ϕ(x)] ( styρ)ϕ(y), ∀ → ∃ ¬ ∧ ∀ ¬ → → ∃ for any internal formula ϕ.   Proof. Immediate by [3, Prop. 3.3 and 5.11].  We will apply underspill most frequently as follows: From ( M Ω)ψ(M) for internal ψ, we conclude that ( K0) st(K) ( M K)ψ(∀M) .∈ Applying ∀ ¬ → ∀ ≥ underspill for ϕ(K) ( M K)ψ(M), we obtain ( stK0)( M K)ψ(M). ≡ ∀ ≥  ∃ ∀ ≥  In conclusion, we have introduced the systems H, P, which are conservative extensions of Peano and Heyting arithmetic with fragments of Nelson’s internal set theory. We have observed that central to the conservation result in Theorem 2.4 is the normal form ( stx)( sty)ϕ(x, y) for internal ϕ. ∀ ∃ NONSTANDARD ANALYSIS AND CONSTRUCTIVISM! 7

2.4. Notations. In this section, we introduce notations relating to H and P. First of all, we mostly use the same notations as in [3]. Remark 2.14 (Notations). We write ( stxτ )Φ(xτ ) and ( stxσ)Ψ(xσ ) as short for ( xτ ) st(xτ ) Φ(xτ ) and ( xσ) st(xσ∀) Ψ(xσ) . We also∃ write ( x0 Ω)Φ(x0) ∀ → ∃ ∧ ∀ ∈ and ( x0 Ω)Ψ(x0) as short for ( x0) st(x0) Φ(x0) and ( x0) st(x0) ∃ ∈   ∀ ¬ → ∃ ¬ ∧ Ψ(x0) . Furthermore, st(x0), is abbreviated by ‘x0 Ω’. A formula A is ‘internal’ ¬  ∈   if it does not involve st, and ‘external’ otherwise. The formula Ast is defined from  A by appending ‘st’ to all quantifiers (except bounded number quantifiers). Secondly, we will use the usual notations for natural, rational and real numbers and functions as introduced in [13, p. 288-289]. We only list the definition of and related notions in P and related systems. Definition 2.15 (Real numbers and related notions in P). (1) A (standard) real number x is a (standard) fast-converging Cauchy sequence 1 0 0 1 q( ), i.e. ( n ,i )( qn qn+i) <0 2n ). We use Kohlenbach’s ‘hat function’ · ∀ | − | from [13, p. 289] to guarantee that every sequence f 1 is a real. 1 1 (2) We write [x](k) := qk for the k-th approximation of a real x = (q( )). · (3) Two reals x, y represented by q( ) and r( ) are equal, denoted ‘x =R y’, if 1 · · n− R ( n)( qn rn 2 1 ). Inequality ‘< ’ is defined similarly. ∀ | − |≤ st 1 (4) We write ‘x y’ if ( n)( qn rn 2n−1 ) and x y if x>y x y. (5) Functions F ≈: R R∀mapping| − reals|≤ to reals are represented≫ by functionals∧ 6≈ 1 1 → Φ → mapping equal reals to equal reals, i.e. ( x, y)(x =R y Φ(x)=R Φ(y)). (RE) ∀ → (6) For a space X with metric X : X R, we write ‘x y’ for ‘ x y X 0’. |·| → ρ 0 ρ ρ≈ 0 | − | ≈ (7) Sets of objects of type ρ are denoted X → , Y →,Z → ,... and are given ρ 0 ρ by their characteristic functions fX→ , i.e. ( x )[x X fX (x) =0 1], ρ 0 ∀ ∈ ↔ where fX→ is assumed to output zero or one. Thirdly, we use the usual extensional notion of equality. Remark 2.16 (Equality). All the above systems include equality between natural numbers ‘=0’ as a primitive. Equality ‘=τ ’ for type τ-objects x, y is defined as:

τ1 τk [x =τ y] ( z ...z )[xz ...zk = yz ...zk] (2.10) ≡ ∀ 1 k 1 0 1 if the type τ is composed as τ (τ1 ... τk 0). Inequality ‘ τ ’ is then just (2.10) with ‘= ’ replaced by ‘ ≡’. In→ the spirit→ of→ Nonstandard Analysis,≤ we define 0 ≤0 ‘approximate equality τ ’ as follows: ≈ st τ1 τk [x τ y] ( z ...z )[xz ...zk = yz ...zk] (2.11) ≈ ≡ ∀ 1 k 1 0 1 with the type τ as above. All the above systems include the axiom of extensionality: ρ ρ ρ τ ( x ,y , ϕ → ) x =ρ y ϕ(x)=τ ϕ(y) . (E) ∀ → However, as noted in [3, p. 1973], the so-called axiom of standard extensionality st (E) is problematic and cannot be included in our systems. We use (E)n+2 to refer to (E) restricted to type n + 2 functionals ϕ. 2.5. Preliminaries. In this section, we introduce the usual defintions of continuity, as well as some fragments of Transfer and Standard Part, their normal forms, and the functionals arising from term extraction as in Theorem 2.4. 8 NONSTANDARD ANALYSIS AND CONSTRUCTIVISM!

2.5.1. Continuity, nonstandard and otherwise. Definition 2.17. [Continuity] A function f is continuous on [0, 1] if 1 1 ( k0)( x [0, 1])( N 0) ( y [0, 1])( x y

Proof. See the proof of Corollary 3.10. 

0 By the previous theorem, term extraction as in Theorem 2.4 converts Π1-TRANS into (µ2), i.e. the former fragment of Transfer is rather non-constructive. Finally, we should point out that it is possible to obtain (some) computational information from (non-constructive) proofs in classical mathematics, even involv- ing comprehension. This is the domain of proof mining, to which Kohlenbach’s monograph [12] provides an excellent introduction. Thus, our use of ‘X is non- constructive’ should be interpreted as the observation that X is rejected in (parts of) constructive mathematics, while X may have implicit constructive content, which can be brought out using proof mining. In fact, we will use a technique from proof theory to obtain computational content from Standard Part in Section 4.2.

2.5.3. Standard Part and related functionals. We shall make use of the following fragments of the Standard Part axiom: ( α1 1)( stβ1 1)(α β), (STP) ∀ ≤1 ∃ ≤1 ≈1 st ( x [0, 1])( y [0, 1])(x y). (STPR) ∀ ∈ ∃ ∈ ≈ These fragments are both equivalent to the following by Theorem 4.4 below: ( T 1 1) ( stn)( β0)( β = n β T ) ( stα1 1)( stn0)(αn T ) (2.18) ∀ ≤1 ∀ ∃ | | ∧ ∈ → ∃ ≤1 ∀ ∈ where ‘T 11’ denotes that T is a binary tree. Clearly, (2.18) is a nonstandard version of ≤weak K¨onig’s lemma, and the latter is a compactness principle by [33, IV].

As it happens, STP and STPR express the nonstandard compactness of Cantor space and the unit interval by Robinson’s theorem in [10, p. 43]. Furthermore, (2.18) is equivalent to the following normal form: ∗ ( stg2)( stw1 )( T 1 1)( (α1 1, k0) w) (αg(α) T ) (2.19) ∀ ∃ ∀ ≤1 ∃ ≤1 ∈ 6∈ ( β 1 1)( i k)(βi T ) → ∀ ≤ ∃ ≤ 6∈ as shown below in Theorem 4.4. Term extraction as in Theorem 2.4 converts (2.19) to the following ‘special fan functional’. Definition 2.21. [Special fan functional] ( g2,T 1 1) ( α1 Θ(g)(2))(α 1 αg(α) T ) (SCF(Θ)) ∀ ≤1 ∀ ∈ ≤1 → 6∈ →  ( β 1 1)( i 0 Θ(g)(1))(βi T ) . ∀ ≤ ∃ ≤ 6∈ ∗ 2 (1 0) Any functional Θ → × satisfying SCF(Θ) is called a special fan functional and the latter object was first introduced in [28]. Note that there is no unique such Θ, i.e. it is in principle incorrect to talk about ‘the’ special fan functional. The computability-theoretic properties of the special fan functional are studied in [20, 28]. Intuitionistically, a special fan functional Θ can be computed (via a term in G¨odel’s T ; see Theorem 4.9) in terms of the intuitionistic fan functional Ω as in MUC(Ω) (See e.g. [13, 3] for the latter) defined as follows: § ( Y 2)( f 1,g1 1)(fΩ(Y )= gΩ(Y ) Y (f)= Y (g)). (MUC(Ω)) ∀ ∀ ≤1 → Classically, Θ can be computed (Kleene’s S1-S9 from [16, 5.1.1]) by ξ as in ( ): § E2 ( ξ3)( Y 2) ( f 1)(Y (f)=0) ξ(Y )=0 . ( ) ∃ ∀ ∃ ↔ E2   10 NONSTANDARD ANALYSIS AND CONSTRUCTIVISM! but Θ cannot be computed (Kleene S1-S9) from any type two functional, which includes rather non-computable functionals like (µ2) and the Suslin functional. ω With regard to first order-strength, RCA0 +( Θ)SCF(Θ) is a conservative extension 2 ∃ of RCA0 + WKL by [13, Prop. 3.15]. All these results may be found in [20]. In conclusion, the special fan functional has rather weak first-order strength, while it is extremely hard to compute (compared to e.g. the computational strength of the Big Five systems of Reverse Mathematics). As will become clear in Sec- tion 4.2, the special fan functional still provides plenty of computational content after applying an extra (algorithmic) step.

3. The intermediate value theorem We discuss Wattenberg’s treatment from [37, II] of the intermediate value theo- rem inside Nonstandard Analysis.

3.1. Introduction and preliminaries. The intermediate value theorem (IVT) is a basic result from and is usually formulated as follows. Theorem 3.1 (IVT). Suppose that f :[a,b] R is continuous and such that → f(a)

Theorem 3.2 (IVT ). Suppose that ε >R 0 and f :[a,b] R is continuous and • → such that f(a)

0 Theorem 3.3 (IVTef (s)). For k and f :[a,b] R uniformly continuous with → 1 modulus g and such that f(a)

3.2. Constructive IVT and Nonstandard Analysis. Wattenberg proves the classical IVT inside Nonstandard Analysis in [37, II.3] using the following steps: b a 0 (i) Define xj := a + jh for h := N− where N is a nonstandard number. (ii) Let j be the largest j such that f(xj ) R 0. 0 ≤ (iii) Since f(xj ) R 0 R f(xj ) and xj xj , we have f(xj ) 0. 0 ≤ ≤ 0+1 0 ≈ 0+1 0 ≈ (iv) Let standard t [0, 1] be such that t xj and conclude f(t )=R 0. 0 ∈ 0 ≈ 0 0 NONSTANDARD ANALYSIS AND CONSTRUCTIVISM! 11

Note that item (iii) makes use of uniform nonstandard continuity, while item (iv) makes use of STP to obtain t0. According to Wattenberg ([37, p. 303]), the final step of the proof is non-constructive (as it involves STP) and should therefore be omitted. Following this approach, the previous steps immediately yield Theorem 3.4 ((1 1) 0) 0 b a where t → × → is defined as follows (with the real h := −N ): (µj N) f jh (2N ) 0 h if such exists t(f,N) := ≤ ≤0 . (3.1) (N +1   otherwise 0 Theorem 3.4 (IVTwat). For nonstandard N and nonstandard uniformly continu- ous f :[a,b] R such that f(a)

We have the following theorem where it should be noted that IVTwat does not have an obvious normal form (to apply Theorem 2.4).

Theorem 3.5. From the proof of IVTwat in P0, a term u can be extracted such that ω RCA0 proves IVTef (u).

Proof. The proof of IVTwat in P0 follows immediately from the definition of t in (4.11) and steps (i)-(iii) above. We now convert IVTwat into a normal form so that we can apply Theorem 2.4; we assume a = 0 and b = 1 for simplicity. Thus, let A(f,k,N) be the conjunction of f(0)

Now bring outside all standard quantifiers as far as possible to obtain: ( stg,l)( f : [0, 1] R)( stn) ( k0)A(f,k,g(k)) ( N n)B(f,N,l) . (3.5) ∀ ∀ → ∃ ∀ → ∀ ≥ Let C(g,l,f,n) be the formula in square brackets in the previous formula. Apply Idealisation I, while bearing in mind Remark 2.10, to push the standard quantifier involving n to the front as follows: ( stg1,l0)( stn0)( f : [0, 1] R)C(g,l,f,n). (3.6) ∀ ∃ ∀ → ω Now apply Theorem 2.4 to ‘P (3.6)’ to obtain a term s such that E-PA ∗ proves ⊢ ( stg1,l0)( n0 s(g,l))( f : [0, 1] R)C(g,l,f,n). (3.7) ∀ ∃ ∈ ∀ → Define u(f,g,k) := t(f, maxi< s(g,k) s(g, k)(i)) and note that IVTef (u).  | | The previous proof serves as a template for obtaining computational content from Nonstandard Analysis as follows. The below proofs follow this template and we will therefore not always go in as much detail as in the previous proof. Template 3.6 (Computational content of Nonstandard Analysis). (i) Bring all sub-formulas into a normal form like in (3.2). (ii) Push all standard quantifiers to the front as follows: (ii.a) If necessary, introduce standard functionals like g in (3.3) and drop ‘st’ in the antecedent like in (3.4). (ii.b) If necessary, use Idealisation (like for (3.5)) to pull standard quantifiers through normal quantifiers. (iii) Obtain a normal form like (3.6) and apply Theorem 2.4 using Remark 2.10.

Furthermore, there is a subtlety involved in the formulation of IVTwat as follows.

Remark 3.7 (On Ω-invariance). By definition, t(f,N) from IVTwat is such that f(t(f,N)) 0 for any N Ω. However, if f has multiple intermediate values, i.e. there are x,≈ y [0, 1] such∈ that f(x) f(y) 0 but x y, then it is possible that t(f,N) t(f,M∈ ) for N,M Ω. In other≈ words,≈ we cannot6≈ use Ω-CA to obtain a standard6≈intermediate value∈ of f. As it turns out, the proof of the theorem also goes through constructively, as follows. This is far from obvious as the proof seems to involve non-constructive steps like the independence of premise principle to go from (3.4) to (3.5).

Corollary 3.8. From the proof of IVTwat in H, a term u can be extracted such that ω E-HA ∗ proves IVTef (u).

Proof. Clearly, the above proof of IVTwat goes through in H. Furthermore, one easily derives (3.7) from IVTwat in H. Indeed, the steps leading up to (3.4) clearly go through in H. For the step from (3.4) to (3.5), the ‘( stl)’ quantifier can be brought to the front in intuitionistic logic, and the same can∀ then be done for the st quantifier ‘( n)’ using the axiom HIP st from Definition 2.11, while bearing in mind Remark∃ 2.10 as usual. Having obtained∀ (3.5), one applies NCR to obtain (3.6), again bearing in mind Remark 2.10. Finally, one applies Theorem 2.12 to ‘H (3.6)’ to obtain the required term.  ⊢ In conclusion, Wattenberg’s claim that Nonstandard Analysis has effective (even constructive) content seems correct in light of Theorem 3.5 and its corollary. NONSTANDARD ANALYSIS AND CONSTRUCTIVISM! 13

3.3. Non-constructivity arising from continuity. In this section, we deal with the exact connection between nonstandard and ε-δ-continuity. We are motivated by the observation that Wattenberg uses the (provable using Transfer) equivalence between nonstandard and ε-δ-continuity without a second thought in [37, III.3]. By the following theorem and Theorem 2.19, any step from ε-δ-continuity to nonstandard continuity, i.e. NSC1 and NSC2 in Definition 2.18, implies a non-trivial fragment of Transfer and is therefore fundamentally non-constructive. 0 Theorem 3.9. The systems P0 + NSC2 and P + NSC1 both prove Π1-TRANS. st Proof. For the first part, we work in P0 + NSC2. Thus, fix standard f0 C ([0, 1]) 0 1 st ∈ and suppose Π1-TRANS, i.e. there is standard h0 such that ( n)h0(n)= 0 ¬ 1 0 ∀ and ( m0)h(m0) = 0. Let b be such that b(q) = q if q = 0 and 1 otherwise. Now define∃ standard6 f as follows: f (x) = f (x) if n0 6 1 (h (n) = 0), 1 1 0 ∀ ≤ b([x](1)) 0 and f0(x) + 1 otherwise. Since for a standard real x [0, 1], the rational [x](n) 0 st ∈  is standard for standard n , we have ( x [0, 1])(f0(x) =R f1(x)), and hence f Cst([0, 1]) by definition. However, f∀ is not∈ nonstandard continuous since: 1 ∈ 1 1 1 f (0) =R f (0) f (0)+1 f ( N )+1=R f ( N ) 1 0 6≈ 0 ≈ 0 2 1 2 for large enough nonstandard N 0. This contradiction finishes the first part. 0 For the second part, we work in P0 + NSC1. Suppose Π1-TRANS, i.e. there is 1 st ¬ standard h0 such that ( n)h0(n)=0 and( m0)h(m0) = 0. Define the standard h(n) ∀ ∃ 6 1 real x as ∞ n . Since 0 x >R 0 the standard function f (x) := is 0 n=0 2 0 2 x +x0 ≈ | | 1 clearly well-defined and continuous (as in f2 C([0, 1])). However, f2(x0)= P 2x0 1 ∈ 6≈ =R f (0) implies that f is not nonstandard continuous. This contradiction x0 2 2 0 yields Π1-TRANS, and we are done.  Recall the definition of ‘modulus-of-continuity-functional’ from Definition 2.18. 0 Corollary 3.10. From the proof that P0 NSC1 Π1-TRANS, a term t can be extracted such that RCAω ( Ξ3) MPC(Ξ)⊢ MU(t→(Ξ))) 0 ⊢ ∀ → 0 Proof. A normal form for Π1-TRANS is given by (2.17), where we use A(f,n) to denote the formula in square brackets. A normal form for nonstandard pointwise continuity (2.13) is obtained as follows. Resolving ‘ ’ in (2.13), we obtain ≈ ( stx [0, 1])( y [0, 1]) ( stN)( x y < 1 ) ( stk)( f(x) f(y) < 1 ) . ∀ ∈ ∀ ∈ ∀ | − | N → ∀ | − | k We may bring out the ‘( stk)’ and ‘( stN)’ quantifiers as follows:  ∀ ∀ ( stx [0, 1])( stk)( y [0, 1])( stN) x y < 1 f(x) f(y) < 1 . ∀ ∈ ∀ ∀ ∈ ∃ | − | N → | − | k ∗ Applying Idealisation I to the underlined formula, we obtain a standard z0 such that ( y [0, 1])( N z) in the previous formula. Now let N be the maximum ∀ ∈ ∃ ∈ 0 of all numbers in z, and note that for N = N0, we have the following: ( stx [0, 1])( stk)( stN)( y [0, 1]) x y < 1 f(x) f(y) < 1 , ∀ ∈ ∀ ∃ ∀ ∈ | − | N → | − | k st st 0 abbreviated by ( x [0, 1], k)( N)B(x,k,N,f ). Hence, NSC1 Π1-TRANS is ∀ ∈ ∃ → ( stg C([0, 1]))( stx [0, 1], k)( stN)B(x,k,N,g) ( stf 1)( stn0)A(f,n), ∀ ∈ ∀ ∈ ∃ → ∀ ∃ which implies (since standard functionals have standard output for standard input): ( stΨ) ( stg C([0, 1]))( stx [0, 1], k)B(x, k, Ψ(x,k,g),g) ( stf 1)( stn0)A(f,n) , ∀ ∀ ∈ ∀ ∈ → ∀ ∃   14 NONSTANDARD ANALYSIS AND CONSTRUCTIVISM! and dropping the ‘st’ in the antecedent, we obtain: ( stΨ) ( g C([0, 1]), x [0, 1], k)B(x, k, Ψ(x,k,g),g) ( stf 1)( stn0)A(f,n) , ∀ ∀ ∈ ∈ → ∀ ∃ and bringing the standard quantifiers up front, we finally have  ( stΨ,f)( stn) ( g C([0, 1]), x [0, 1], k)B(x, k, Ψ(x,k,g),g) A(f,n) . (3.8) ∀ ∃ ∀ ∈ ∈ → Applying Theorem 2.4 to ‘P (3.8)’, we obtain a term t such that  0 ⊢ ( Ψ,f)( n t(Ψ,f)) ( g C([0, 1]), x [0, 1], k)B(x, k, Ψ(x,k,g),g) A(f,n) . ∀ ∃ ∈ ∀ ∈ ∈ → which is exactly as required in light of the definition of A, B and Remark 2.10. 

By the previous theorem and corollary, NSC1 translates to the existence of a modulus-of-continuity-functional when applying Theorem 2.4. Such a functional is fundamentally non-constructive by the corollary, and this non-constructiveness ‘trickles down’ to any nonstandard theorem of P0 + NSC1 as follows.

st st Corollary 3.11. Let ϕ be internal. From P0 + NSC1 ( x)( y)ϕ(x, y), a term t can be extracted such that RCAω ( Ξ3) MPC(Ξ) ⊢( ∀x)( y∃ t(x, Ξ))ϕ(x, y) . 0 ⊢ ∀ → ∀ ∃ ∈ Proof. Analogous to the proof of the previous corollary following Template 3.6 

The previous theorem and corollaries imply that the step from ε-δ continuity (relative to ‘st’ or not) to the nonstandard variety always involves a non-trivial instance of Transfer, which is fundamentally non-constructive. In particular, by Corollary 3.11, any result proved using NSC1 only provides computational infor- mation involving a non-constructive modulus-of-continuity-functional. In general, moving from the standard into the nonstandard world is highly non-constructive (requiring Transfer), as was sketched in Section 1 in the form of Osswald’s local constructivity. Nonetheless, Wattenberg freely uses Transfer and the equivalence between ‘ε-δ’ and nonstandard continuity in [37, II-III]. This aspect of his investiga- tion into the computational content of Nonstandard Analysis thus seems incorrect. Furthermore, as suggested by its proof, Theorem 3.9 goes through for other no- tions besides continuity. We now show, for differentiability and Riemann integra- 0 tion, that the step from the ε-δ definition to the nonstandard one yields Π1-TRANS. First of all, we have the usual definition of differentiability. Definition 3.12. A function f is nonstandard differentiable at a if

′ f(a+ε) f(a) f(a+ε ) f(a) ( ε,ε′ = 0) ε,ε′ 0 − ′− . (3.9) ∀ 6 ≈ → ε ≈ ε A function f is differentiable at a if  ′ 0 0 1 f(a+ε) f(a) f(a+ε ) f(a) 1 ( k )( N )( ε,ε′) 0 < ε , ε′ < − ′− < . (3.10) ∀ ∃ ∀ | | | | N → ε − ε k A ‘modulus of differentiability at a’ is a function g1 such that g(k) is N 0 in (3.10).

Let NSD be the statement any standard f : R R differentiable at zero is also nonstandard differentiable there. Now, NSD is a→ theorem of IST but we also have the following implication.

0 Theorem 3.13. The system P + NSD proves Π1-TRANS. NONSTANDARD ANALYSIS AND CONSTRUCTIVISM! 15

0 1 Proof. Working in P+NSD, suppose we have Π1-TRANS, i.e. there is standard h0 st ¬ such that ( n)h0(n)=0 and ( m0)h(m0) = 0. Define the standard real x0 as in ∀ ∃ 6 1 x2+x the proof of Theorem 3.9. Since 0 x0 >R 0 the standard function f0(x) := e 0 is well-defined and differentiable in≈ the usual internal ‘epsilon-delta’ sense. However,

1 1 1 1 1 1 2x x 1 x 1 x 2x x f0(√x0) f0(0) e 0 e 0 x e 0 1 x e 0 1 e 0 e 0 f0( √x0) f0(0) − = − = e 0 − 0 e 0 − = − = − − √x0 √x0 √x0 √x0 √x0 √x0 ≫ ≫ − − − which implies that f0 is not nonstandard differentiable at zero. This contradiction 0 yields Π1-TRANS, and we are done.  Let DIF(Ξ) be the statement that Ξ(f) is a modulus for differentiability at zero for every f differentiable at zero.

0 Corollary 3.14. From the proof that P NSD Π1-TRANS, a term t can be extracted such that E-PAω ( Ξ3) DIF(Ξ) ⊢ MU(t→(Ξ))) ⊢ ∀ → Proof. A normal form for differentiability as in (3.9) is easy to obtain and as follows:

′ st 0 st 0 1 f(a+ε) f(a) f(a+ε ) f(a) 1 ( k )( N )( ε,ε′ = 0) ε , ε′ < − ′− < , ∀ ∃ ∀ 6 | | | | N → ε − ε k

The proof is straightforward and analogous to the proof of Corollary 3.10.  

Hence, switching from epsilon-delta differentiability to the nonstandard variety 2 as in NSD is at least as non-constructive as (µ ) and NSC1. One readily obtains a version of NSC Π0-TRANS for NSD, i.e. for ε-δ-differentiability relative to ‘st’. 2 → 1 Next, we consider the usual definitions of Riemann integration. Definition 3.15. [Riemann Integration]

(1) A partition of [0, 1] is an increasing sequence π = (0,t0, x1,t1,...,xM 1,tM 1, 1). We write ‘π P ([0, 1])’ to denote that π is such a partition. − − (2) For π P ([0∈, 1]), π is the mesh, i.e. the largest distance between two ∈ k k adjacent partition points xi and xi+1. R R M 1 (3) For π P ([0, 1]) and f : , the real Sπ(f) := i=0− f(ti)(xi+1 xi) is the ∈Riemann sum of f and→ π. − (4) A function f is nonstandard integrable on [0, 1] if P

( π, π′ P ([0, 1])) π , π′ 0 Sπ(f) Sπ′ (f) . (3.11) ∀ ∈ k k k k≈ → ≈ (5) A function f is integrableon [0, 1] if  0 0 1 1 ( k )( N )( π,ρ P ([0, 1])) π , ρ < Sπ(f) Sρ(f) < . (3.12) ∀ ∃ ∀ ∈ k k k k N → | − | k A modulus of (Riemann) integration ω1 provides N = ω(k) as in (3.12). Let NSR be the statement a standard f : R R integrable on the unit interval is also nonstandard integrable there. As above,→NSR is a theorem of IST but we also have the following implication.

0 Theorem 3.16. The system P + NSR proves Π1-TRANS. 0 Proof. Suppose NSR Π1-TRANS and note that f0 from Theorem 3.13 is Riemann ∧¬ 1 1 x 2x integrable. However, since the distance between f0(0) = e 0 and f0(√x0) = e 0 is larger than any standard real, replacing 0 by √x0 in a partition causes the associated Riemann sums to be apart by more than an infinitesimal.  16 NONSTANDARD ANALYSIS AND CONSTRUCTIVISM!

Let RIE(κ) be the statement that κ(f) is a modulus of Riemann integration for every f : R R integrable on the unit interval. → 0 Corollary 3.17. From the proof that P NSR Π1-TRANS, a term t can be extracted such that E-PAω ( κ3) RIE(κ) ⊢ MU(t→(κ))) ⊢ ∀ → Proof. The proof is straightforward and analogous to that of Corollary 3.10. 

Hence, switching from epsilon-delta integrability to the nonstandard variety as 2 in NSR is at least as non-constructive as (µ ) and NSC1. One readily obtains a version of NSC Π0-TRANS for NSR, i.e. for ε-δ-integrability relative to ‘st’. 2 → 1 In conclusion, Wattenberg’s claim that Nonstandard Analysis has effective (even constructive) content is correct in light of Theorem 3.5; his implementation using ε-δ continuity and Transfer is problematic in light of Theorem 3.9, but easily sal- vageable: By Theorem 3.5 it suffices to just adopt nonstandard (rather than ε-δ) continuity, in line with Osswald’s local constructivity. We next investigate Watten- berg’s claims regarding the constructive status of Standard Part in Section 3.4.

3.4. The non-constructive status of Standard Part. We investigate the con- structive status of Standard Part in light of Wattenberg’s claims that it be funda- mentally non-constructive. First of all, we have the following theorem regarding the use of STP, which is the only fragment of Standard Part used by Wattenberg in [37].

st st Theorem 3.18. Let ϕ be internal. From P0 + STP ( x)( y)ϕ(x, y), a term t can be extracted such that RCAω ( Θ3) SCF(Θ) ⊢( ∀x)( y ∃ t(x, Θ))ϕ(x, y) . 0 ⊢ ∀ → ∀ ∃ ∈ Proof. Note that STP is equivalent to (2.19) form Section 2.5.3 by Theorem 4.4. The theorem now follows easily by following the proof of Corollary 3.10. 

In light of the previous theorem, the use of STP in the proof of a nonstandard theorem translates to the presence of the special fan functional Θ after applying Theorem 2.4. Given the computational hardness of Θ, Wattenberg’s claims regard- ing the non-constructive nature of Standard Part seem justified. However, as will be established in Section 4.2, the fragment of Standard Part used by Wattenberg (namely STP from Section 2.5.3) does have plenty of effective content, though extra technical machinery is needed for this. Secondly, we show that the generalisation of STP to type two functionals is non- constructive as it implies ( 2). In particular, the following rather weak fragment of Standard part is established∃ to be non-constructive by Theorem 3.19. ( Y 2 1)( stZ2 1)(Z Y ). (STP ) ∀ ≤2 ∃ ≤2 ≈2 2 st Note that (E)n+2 results in a conservative extension of P0 as shown in [4]. Theorem 3.19. The system P + (E)st + STP proves ( 2)st. 0 2 2 ∃ st Proof. In a nutshell, we work relative to ‘st’ in P0 + (E)2 + STP2 and define a 0 functional which computes the separating set in Σ1-separation ([33, I.11.7]). The 0 theorem then follows from the equivalence between the uniform version of Σ1- separation and ( 2), as proved in [27, Theorem 3.6]. ∃ NONSTANDARD ANALYSIS AND CONSTRUCTIVISM! 17

0 Let f1,f2 be standard binary sequences and fix nonstandard N . Let K(n,f1,f2) 0 be the largest k N such that ( n1,n2 k)(f1(n1,n) = 0 f2(n2,n) = 0), if such number exists,≤ and zero otherwise.∀ Define≤ Y 2 as follows:6 ∨ 6

1 ( n1 K(n,f1,f2))(f1(n1,n)=0) Y (f1,f2,n) := ∃ ≤ . (0 otherwise st 0 st st Now suppose ( n )( ϕ1 (n) ϕ2 (n)) where ϕi(n) ( ni)(f(ni,n) = 0). By ∀ ¬ ∨ ¬ 0≡ ∃ overspill, K(n,f1,f2) is nonstandard for all standard n . By definition, we have ( stn0) ϕst(n) Y (f ,f ,n)=1 ϕst(n) Y (f ,f ,n)=0 . ∀ 1 → 1 2 ∧ 2 → 1 2 2 Now apply STP2to obtain standard Z such that Z 2 Y . Then Z(f1,f2,n) is 0 ≈ standard and provides the separating set from Σ1-separation for standard inputs and relative to ‘st’.  We provide an alternative proof for Theorem 3.19 as follows. Proof. We prove that STP UWKLst in P, where UWKL is as follows: 2 → 1 1 1 0 0 0 ( Φ → )( T 1) ( n )( β )(β T β = n) ( m )(Φ(T )m T ) . (3.13) ∃ ∀ ≤1 ∀ ∃ ∈ ∧ | | → ∀ ∈ As proved in [15], UWKL implies ( 2), and the latter proof immediately transfers to P + (E)st, yielding that STP ∃ ( 2)st. Fix a standard binary tree T . 0 2 2 → ∃ Apply overspill to ( stn0)( β0)(β T β = n) to obtain a sequence in T of nonstandard length, say∀ N. Now∃ define∈ Φ(∧T )(0) | | as 0 (resp. 1) if there is a sequence ∗ β0 T of length N such that β(0) = 0 (resp. if this is not the case). Then define ∈ ∗ Φ(T )(n +1) as Φ(T )(0) ... Φ(T )(n) 0 (resp. 1) if there is a sequence β0 T ∗ ∗ ···∗ ∈ of length N such that Φ(T )(0) ... Φ(T )(n) 0= β(n + 1) (resp. if this is not the ∗ ∗ 1 case). By STP2, there is standard Ψ such that Φ(f) 1 Ψ(f) for standard f 1 1, st ≈ ≤ and STP UWKL follows immediately.  2 → In conclusion, the previous theorem suggests that the axiom Standard Part is in general fundamentally non-constructive, as claimed by Wattenberg. Moreover, st since (E)n+2 readily follows from Transfer, Theorem 3.19 is especially relevant when a proof utilises both Transfer and Standard Part.

4. Compactness We discuss Wattenberg’s treatment from [37, III] of compactness (Sections 4.1 and 4.2) and the associated extreme value theorem (Section 4.3). 4.1. Constructive compactness and Nonstandard Analysis. Wattenberg de- scribes the following form of compactness as ‘acceptable’ in [37, III.4]. As we will see, his choice of compactness is indeed most suitable for obtaining constructive or effective results. Definition 4.1. [F -compactness] A metric space X is F -compact if there is a 0 standard sequence x( ) such that ( x X)( N Ω)( k N)( xk x X 0). · ∀ ∈ ∀ ∈ ∃ ≤ | − | ≈ Note that inside P (and extensions), the unit interval and Cantor space are F - 0 compact, but not necessarily nonstandard compact (as P+Π1-TRANS STP by [20, 4]). In particular, nonstandard compactness guarantees the infinitesimal6⊢ proximity of§ a standard point, while F -compactness states the presence of an ‘infinitesimal grid’ of nonstandard points. Thus, F -compactness expresses the intuitive notion 18 NONSTANDARD ANALYSIS AND CONSTRUCTIVISM! that a compact space ‘can be divided into infinitesimal pieces’, a mainstay of the infinitesimal calculus used in physics and engineering. The notion of F -compactness for special cases has been studied in [29]. We first prove a basic result regarding F -compactness. Note that the latter provides a kind of ‘discretisation’ of the space X as used in an essential way for the unit interval in the steps (i)-(iv) at the beginning of Section 3.2.

Theorem 4.2 (FCR). An F -compact X R has a supremum, i.e. for all standard ⊂ x( ) and any X R, we have · ⊂ 0 ( x X)( N Ω)( k N)(xk x) ( x X)( N Ω)(x / t(x( ),N))), ∀ ∈ ∀ ∈ ∃ ≤ ≈ → ∀ ∈ ∀ ∈ · where t(x( ),N) := maxi N xi. · ≤ Note that Ω-CA converts t(x( ),N) from the theorem into a standard supremum. · The constructive version of the previous theorem is [5, Theorem 3, p. 34]. The latter version involves the notion of ‘totally boundedness’ as in the antecedent of (4.1), which ‘falls out’ of the notion of F -compactness by the following theorem. ω Theorem 4.3. From a proof P0 FCR, a term s can be extracted such RCA0 proves ⊢ 1 that for any x( ) and X R and g , we have · ⊂ 0 1 ( k , x X)( n g(k))( xn x

Theorem 4.4. In P, STP is equivalent to STPR and to the normal forms ∗ ( stg2)( stw1 , k0)( T 1 1) ( α1 w)(αg(α) T ) (4.2) ∀ ∃ ∀ ≤1 ∀ ∈ 6∈  ( β 1 1)( i k)(βi T ) , → ∀ ≤ ∃ ≤ 6∈ ∗ 1  ( stg2)( stw1 , k)( z R) y (w [0, 1]) ( y z >R ) (4.3) ∀ ∃ ∀ ∈ ∀ ∈ ∩ | − | g(y) 1   ( x [0, 1])( x z >R ) . → ∀ ∈ | − | k  NONSTANDARD ANALYSIS AND CONSTRUCTIVISM! 19

Proof. We first prove that STP and STPR are equivalent. Now, Hirst establishes in [8] that RCA0 proves that every real x [0, 1] has a binary expansion, i.e. 1 α(i) ∈ ( x [0, 1])( α 1)(x =R ∞ i ). Since P proves the latter (both the ∀ ∈ ∃ ≤1 i=0 2 0 internal version and the version relative to ‘st’), it is clear that STP STPR. P ↔ Secondly, we prove that STP is equivalent to

( T 1 1) ( stn)( β0)( β = n β T ) (4.4) ∀ ≤1 ∀ ∃ | | ∧ ∈ ( stα1 1)( stn0)(αn T ) .  → ∃ ≤1 ∀ ∈ st 0  0 Assume STP and apply overspill to ( n)( β )( β = n β T ) to obtain β0 T ∀ ∃ | | 1 ∧ ∈ ∈ with nonstandard length β0 . Now apply STP to β := β0 00 ... to obtain a 1 | | st ∗ standard α 1 1 such that α 1 β and hence ( n)(αn T ). For the reverse ≤1 ≈ ∀ ∈ direction, let f be a binary sequence, and define a binary tree Tf which contains all initial segments of f. Now apply (4.4) for T = Tf to obtain STP.

Thirdly, assume STP and note that the contraposition of (4.4) yields:

( T 1 1) ( stα 1)( stn0)(αn T ) (4.5) ∀ ≤1 ∀ ≤1 ∃ 6∈ → st 0  ( k )( β 1 1)( i k)(βi T ) . ∃ ∀ ≤ ∃ ≤ 6∈ Since standard functionals have standard output for standard input, (4.5) implies:

( T 1 1)( stg2) ( stα 1)(αg(α) T ) (4.6) ∀ ≤1 ∀ ∀ ≤1 6∈ → st 0  ( k )( β 1 1)( i k)(βi T ) . ∃ ∀ ≤ ∃ ≤ 6∈ Pushing all standard quantifiers outside as far as possible, we obtain that 

( stg2)( T 1 1)( stk0, α1 1) (αg(α) T ) (4.7) ∀ ∀ ≤1 ∃ ≤1 6∈ ( β 1 1)( i k)(βi T ) . → ∀ ≤ ∃ ≤ 6∈ Applying Idealisation I, we pull the standard quantifiers to the front as follows:

∗ ( stg2)( stw1 )( T 1 1)( (α1 1, k0) w) (αg(α) T ) (4.8) ∀ ∃ ∀ ≤1 ∃ ≤1 ∈ 6∈  ( β 1 1)( i k)(βi T ) . → ∀ ≤ ∃ ≤ 6∈ Now assume (4.8) and note that since w is standard, all of its elements are, implying (4.7). Bringing all standard quantifiers inside again (as far as possible), we obtain st st 0 (4.6). We now immediately obtain (4.5) by noting that ( α 1 1)( n )(αn ∗ st 1 0 st 0 ∀ ≤ ∃ 6∈ T ) implies ( Φ → )( α 1)( n Φ(α))(αn T ) by applying HAC and ∃ ∀ ≤1 ∃ ∈ 6∈ int defining g(α) := maxi< Φ(α) Φ(α)(i) as in Remark 2.10.  | |

In conclusion, STP is equivalent to the normal form (4.2), and term extraction as in Theorem 2.4 converts the latter to the special fan functional introduced in Section 2.5. Since the latter boasts extreme computational hardness, it indeed seems better to avoid nonstandard compactness STP in favour of F -compactness. In the next section, we will show how computational content can still be obtained from STP and the special fan functional. 20 NONSTANDARD ANALYSIS AND CONSTRUCTIVISM!

4.2. The constructive status of nonstandard compactness. We investigate Standard Part in light of Wattenberg’s claims that it be fundamentally non-constructive. By Theorem 3.19, a rather small fragment of Standard part is indeed fundamen- tally non-constructive. By contrast, we show in this section that STP, which is the particular fragment Wattenberg uses in [37, II-III], still yields constructive content, after some extra technical steps. This section is somewhat more technical in nature as we assume familiarity with the ECF-translation from [35, 2.6.5]. § First of all, nonstandard continuity (2.13) clearly yields uniform nonstandard continuity (2.15) for the unit interval inside P0 + STP by Theorem 4.4. Thus, the following version of IVT is immediate where the term t is as in (4.11).

0 Theorem 4.5 (IVTwat′ ). For nonstandard N and nonstandard continuous f : [0, 1] R such that f(0)

Theorem 4.7. From the proof of IVTwat′ in P0 + STP, a term u can be extracted ω 3 such that RCA proves ( Θ ) SCF(Θ) IVT′ (u(Θ)) . 0 ∀ → ef

Proof. Since P0 proves IVTwat by Theorem 3.19, P0+ STP proves IVTwat′ , as in the latter system every nonstandard continuous function is automatically uniform nonstandard continuous on the unit interval by STPR and Theorem 4.4. To obtain the effective results from the theorem, one just proceeds as in Theorem 3.5 using the normal form (4.2) of STP. 

Secondly, Theorem 4.7 is not very satisfactory as the special fan functional is not computable (in the sense of Kleene’s S1-S9) in e.g. ( 2), or even the Suslin functional (See [20, 3] for these results). However, the following∃ observation ([13, 2]) by Kohlenbach§ will be seen to solve this problem in Theorem 4.9: § If RCAω A then RCA2 [A] . (4.9) 0 ⊢ 0 ⊢ ECF 2 Here, RCA0 is essentially the base theory RCA0 of Reverse Mathematics ([33, II]) formulated with function rather than set variables; the syntactical interpretation [ ]ECF is defined in [35, 2.6.5] and is based on the Kleene-Kreisel model of con- tinuous· functionals. In the§ latter, higher-type objects are represented by so-called associates which is equivalent to the representation used in Reverse Mathematics by [14, Prop. 4.4] of continuous functionals on Baire space. In a nutshell, the ECF-translation amounts to replacing all objects of type two or higher by type one associates. Applying ECF as in (4.9) to the final part of Theorem 4.7, we shall observe that the special fan functional is converted into a ‘more computable’ object. We now introduce the definition of associate for a type two functional from [14], and study the intuitionistic fan functional as an example. Definition 4.8. [Associate] The function α1 is an associate for continuous Φ2 if (1) ( β1)( n0)(α(βn) > 0), ∀ ∃ (2) ( β1,m0)(α(βm) > 0 Φ(β)+1= α(βm)). ∀ → NONSTANDARD ANALYSIS AND CONSTRUCTIVISM! 21

One often writes α(β), to be understood as α(βm) 1 for large enough m as in the first item. Given an associate α1 for Φ2, an associate− γ1 for Ψ3 is now defined such that Ψ(Φ) is γ(α) where the latter is again γ(αk) 1 for large enough k. − By way of an example, consider the intuitionistic fan functional as in MUC(Ω) from Section 2.5.3. Following the heuristic that all objects of type two or higher are replaced by associates by ECF, it is straightforward to see that [MUC(Ω)]ECF is:

( α1) ( f 1)( n0)(α(fn) > 0) ( m0)(γ(αm) > 0) (TOF(γ1)) ∀ ∀ ∃ → ∃ ∧ h ( g1,h1 1, k0)([γ(αk) >0 hγ(αk)= gγ(αk)] α(h)= α(g) > 0) . ∀ ≤1 ∧ 0 → i The underlined formula expresses that α1 is an associate representing a (continuous) functional Y 2, while TOF(γ) expresses that γ1 is an associate for the intuitionistic fan functional, i.e. Ω(Y ) as in MUC(Ω) is given by γ(α), and the rest of TOF(γ) ensures that γ(α) makes sense. Note that if a functional Φ2 has an associate as in Definition 4.8, it is automat- ically continuous on Baire space. Thus, since µ2 as in (µ2) is discontinuous (e.g. 2 at 00 ... ), [(µ )]ECF is equivalent to 0 = 1. In particular, we observe that the ECF- translation replaces any type two variable with a type one variable over associates, i.e. the new variable ranges over (representations of) continuous functionals. As a result of the aforementioned ‘continuous replacement’, WKL is equivalent ( γ1)TOF(γ), and the latter is of course [( Ω3)MUC(Ω)] . As it turns out, the ∃ ∃ ECF intuitionistic fan functional even has a primitive recursive associate FAN which may be found in [19, p. 102]. We thus also have WKL TOF(FAN) (See [16, 7.3.4]). The following theorem shows that the special↔ fan functionaldbecomes § ‘more computable’ thanks to ECF. We shall make use of the nonstandardd axiom ( stY 2)( f 1,g1 1)(f g Y (f)= Y (g)), (NUC) ∀ ∀ ≤1 ≈1 → 0 Note that NUC expresses that every type two functional is nonstandard uniformly continuous on Cantor space, akin to Brouwer’s continuity theorem ([6]). Theorem 4.9. From P NUC STP, terms t ,t can be extracted such that 0 ⊢ → 0 1 RCAω ( Ω)(MUC(Ω) SCF(t (Ω))) and RCA2 + WKL [SCF] (t (FAN)). 0 ⊢ ∀ → 0 0 ⊢ ECF 1 Proof. As in Theorem 3.5, the normal form of NUC is readily obtained as follows: d ( stY 2)( stN)( f 1,g1 1)(fN = gN Y (f)= Y (g)), (4.10) ∀ ∃ ∀ ≤1 → 0 Applying HACint to (4.10) as in Remark 2.10 yields that ( stΩ3)( stY 2)( stN)( f 1,g1 1)(fΩ(Y )= gΩ(Y ) Y (f)= Y (g)). ∃ ∀ ∃ ∀ ≤1 → 0 To obtain the nonstandard implication, note that STP is equivalent to (4.2) and define k0 as in the latter as the (clearly standard) maximum of g(σ 00) for all ∗ 0 1 ∗ binary σ of length Ω(g) + 1, while w is the (standard) collection of all σi 00 ... ∗ 0 ∗ where the binary σi has length Ω(g)+1. Hence, NUC implies (4.2) and hence STP. Since NUC and STP have normal forms (4.10) and (4.2), applying term extraction as in Theorem 2.4 to P [NUC STP] readily yields the term t from the theorem. 0 ⊢ → 0 Finally, applying the ECF-translation as in (4.9) to RCAω ( Ω)(MUC(Ω) 0 ⊢ ∀ → SCF(t (Ω))), we obtain that RCA2 [( Ω)(MUC(Ω) SCF(t (Ω)))] , which 0 0 ⊢ ∀ → 0 ECF 22 NONSTANDARD ANALYSIS AND CONSTRUCTIVISM! becomes RCA2 ( γ1)(TOF(γ) [SCF] (t (γ))) as [MUC(Ω)] is TOF(γ). 0 ⊢ ∀ → ECF 1 ECF Since WKL TOF(FAN) by [16, 7.3.4], the theorem now follows.  ↔ § Note that parts ofd Theorem 4.9 may be found in [20,28]. We have the following corollary pertaining to Theorem 4.7.

1 Corollary 4.10. From the proof of IVTwat′ in P0 + STP, a term v can be extracted 2 such that RCA0 + WKL proves [IVTef′ ]ECF(v(FAN)) . Proof. Apply the ECF-translation to the conclusion of Theorem 4.7.  d Note that WKL is non-constructive, but the term v(FAN) is computable. Fur- thermore, the only real modification the ECF-translation bestows upong the inter- mediate value theorem from Theorem 4.7 is the replacementd of continuous functions by associates (which can always be done given WKL by [14, Theorem 4.6]). In conclusion, we have observed that STP is indeed non-constructive in nature in that it gives rise to the special fan functional as in Theorem 4.7. However, a somewhat technical detour (using the ECF-interpretation) still yields computational information as in Corollary 4.10.

4.3. Extreme value theorem. We briefly discuss Wattenberg’s treatment from [37, III] of Weierstraß’ extreme value theorem inside Nonstandard Analysis.

4.3.1. Preliminaries. The extreme value theorem (WMX) is a basic result from calculus and can be formulated as follows. Theorem 4.11 (WMX). Suppose that X is compact and that f : X R is con- → tinuous. Then there is x X such that ( y X)(f(y) R f(x)). ∈ ∀ ∈ ≤ As is well-known, WMX implies a non-trivial fragment of the law of excluded middle (See e.g. [2, I.6] or [17]) and is therefore rejected in constructive mathemat- ics. A slight modification of Wattenberg’s nonstandard version of WMX will be shown to yield the following effective version in Section 4.3.2, similar to [5, p. 89].

0 Theorem 4.12 (WMXef (s)). For k and f : X R uniformly continuous with modulus g on the compact space X with modulus of→ totally boundedness h, we have 1 ( x X)(f(x) R f(s(f,g,h,k)) + ). ∀ ∈ ≤ k Note that this version no longer involves Nonstandard Analysis. Furthermore, the term s is ‘read off’ from the (modified) Wattenberg proof. Thus, Wattenberg’s claims about the effective content of Nonstandard Analysis are again at least par- tially correct. One the other hand, Wattenberg explicitly uses Transfer in the proof of [37, III.6], which is problematic if one is interested in computational content, as was established above.

4.3.2. Constructive extreme value theorem and Nonstandard Analysis. Wattenberg proves various nonstandard versions of WMX inside Nonstandard Analysis in [37, III.7]. He refers to (a trivial reformulation of) the Theorem 4.13 below as a completely “constructive” version of the Extreme Value Theorem in [37, p. 308]. Note that Wattenberg uses the notion of an ‘implementation’ (See [37, Def. III.8]) rather than nonstandard uniform continuity, although both essen- tially amount to the same thing in this context. NONSTANDARD ANALYSIS AND CONSTRUCTIVISM! 23

0 Theorem 4.13 (WMXwat). For nonstandard N , F -compact X with standard se- quence x( ), and nonstandard uniformly continuous f : X R, we have ( x · → ∀ ∈ X)(f(x) / f(t(x( ),N))). · ((1 1) 0) 0 b a The term t → × → from the theorem is defined as follows (where h := −N ): (µj N) f jh (2N ) 0 h if such exists t(f,N) := ≤ ≤0 . (4.11) (N +1   otherwise Wattenberg proves WMXwat in [37, III.7] using (what amounts to) the following: 0 (i) Fix a standard sequence x( ) and nonstandard N as provided by the F - compactness of X. · (ii) Define t0 := x0 and tn+1 := tn if f(xn+1) / f(tn) and tn+1 = xn+1 if f(xn ) ' f(tn). Note that for n N, we have f(xn) / f(tN ). +1 ≤ (iii) Since for every x X there is j N such that xj x, we have ( x ∈ ≤ ≈ ∀ ∈ X)(f(x) / f(tN )) by continuity. Note that item (iii) makes use of uniform nonstandard continuity. When working in IST, one would apply STP to tN to obtain a standard maximum for f. Note that a similar remark regarding Ω-CA as in Remark 3.7 applies to WMXwat. Finally, we have the following theorem.

Theorem 4.14. From the proof of WMXwat in P0, a term u can be extracted such ω that RCA0 proves WMXef (u).

Proof. The proof of WMXwat inside P0 follows from the above steps (i)-(iii), assum- N ing we use approximations (say up to precision 2 for N from item (i)) to f(xi) and f(ti). One then applies Theorem 2.4 to P WMX to obtain the term u 0 ⊢ wat from the theorem, following Template 3.6.  References

[1] Sergio Albeverio, Raphael Høegh-Krohn, Jens Erik Fenstad, and Tom Lindstrøm, Nonstan- dard methods in stochastic analysis and mathematical physics, Pure and Applied Mathemat- ics, vol. 122, Academic Press, 1986. [2] Michael J. Beeson, Foundations of constructive mathematics, Ergebnisse der Mathematik und ihrer Grenzgebiete, vol. 6, Springer, 1985. Metamathematical studies. [3] Benno van den Berg, Eyvind Briseid, and Pavol Safarik, A functional interpretation for nonstandard arithmetic, Ann. Pure Appl. Logic 163 (2012), no. 12, 1962–1994. [4] Benno van den Berg and Sam Sanders, Transfer equals Comprehension, Submitted (2014). Available on arXiv: http://arxiv.org/abs/1409.6881. [5] Errett Bishop, Foundations of constructive analysis, McGraw-Hill Book Co., New York, 1967. [6] L. E. J. Brouwer, Collected works. Vol. 1, North-Holland Publishing Co., Amsterdam, 1975. Philosophy and foundations of mathematics; Edited by A. Heyting. [7] Damir D. Dzhafarov, Reverse Mathematics Zoo. http://rmzoo.uconn.edu/. [8] Jeffry L. Hirst, Representations of reals in reverse mathematics, Bull. Pol. Acad. Sci. Math. 55 (2007), no. 4, 303–316. [9] Anton Jensen, A computer oriented version of “non-standard analysis”, Contributions to non-standard analysis (Sympos., Oberwolfach, 1970), North-Holland, 1972, pp. 281–289. Studies in Logic and Found. Math., Vol. 69. [10] Albert E. Hurd and Peter A. Loeb, An introduction to nonstandard real analysis, Pure and Applied Mathematics, vol. 118, Academic Press Inc., Orlando, FL, 1985. [11] H. Jerome Keisler, The hyperreal line, Real numbers, generalizations of the reals, and theories of continua, Synthese Lib., vol. 242, Kluwer Acad. Publ., Dordrecht, 1994, pp. 207–237. [12] Ulrich Kohlenbach, Applied proof theory: proof interpretations and their use in mathematics, Springer Monographs in Mathematics, Springer-Verlag, Berlin, 2008. 24 NONSTANDARD ANALYSIS AND CONSTRUCTIVISM!

[13] , Higher order reverse mathematics, Reverse mathematics 2001, Lect. Notes Log., vol. 21, ASL, 2005, pp. 281–295. [14] , Foundational and mathematical uses of higher types, Reflections on the foundations of mathematics (Stanford, CA, 1998), Lect. Notes Log., vol. 15, ASL, 2002, pp. 92–116. [15] , On uniform weak K¨onig’s lemma, Ann. Pure Appl. Logic 114 (2002), no. 1-3, 103– 116. Commemorative Symposium Dedicated to Anne S. Troelstra (Noordwijkerhout, 1999). [16] John Longley and Dag Normann, Higher-order Computability, Theory and Applications of Computability, Springer, 2015. [17] Mark Mandelkern, Constructive continuity, Mem. Amer. Math. Soc. 42 (1983), v+117. [18] Edward Nelson, Internal set theory: a new approach to nonstandard analysis, Bull. Amer. Math. Soc. 83 (1977), no. 6, 1165–1198. [19] Dag Normann, Recursion on the countable functionals, LNM 811, vol. 811, Springer, 1980. [20] Dag Normann and Sam Sanders, Nonstandard Analysis, Computability Theory, and their connections, Submitted, Available from arXiv: https://arxiv.org/abs/1702.06556 (2017). [21] Horst Osswald, Computation of the kernels of L´evy functionals and applications, Illinois Journal of Mathematics 55 (2011), no. 3, 815–833. [22] , Malliavin calculus for L´evy processes and infinite-dimensional Brownian motion, Cambridge Tracts in Mathematics, vol. 191, Cambridge University Press, Cambridge, 2012. [23] Erik Palmgren, Constructive nonstandard analysis, M´ethodes et analyse non standard, Cahiers Centre Logique, vol. 9, Acad.-Bruylant, Louvain-la-Neuve, 1996, pp. 69–97. [24] David A. Ross, A nonstandard proof of a lemma from constructive measure theory, MLQ Math. Log. Q. 52 (2006), no. 5, 494–497. [25] , The constructive content of nonstandard measure existence proofs—is there any?, Reuniting the antipodes—constructive and nonstandard views of the continuum (Venice, 1999), Synthese Lib., vol. 306, Kluwer, 2001, pp. 229–239. [26] , Non-standard analysis, North-Holland, Amsterdam, 1966. [27] Nobuyuki Sakamoto and Takeshi Yamazaki, Uniform versions of some axioms of second order arithmetic, MLQ Math. Log. Q. 50 (2004), no. 6, 587–593. [28] Sam Sanders, The Gandy-Hyland functional and the computational aspect of Nonstandard Analysis, To appear in Computability, http://arxiv.org/abs/1502.03622 (2017). [29] , The unreasonable effectiveness of Nonstandard Analysis, Submitted; Available from arXiv: http://arxiv.org/abs/1508.07434 (2015). [30] , The taming of the Reverse Mathematics zoo, Submitted, http://arxiv.org/abs/1412.2022 (2015). [31] , The refining of the taming of the Reverse Mathematics zoo, To appear in Notre Dame Journal for Formal Logic, http://arxiv.org/abs/1602.02270 (2016). [32] Peter Schuster, Ulrich Berger, and Horst Osswald (eds.), Reuniting the antipodes: constructive and nonstandard views of the continuum, Synthese Library, vol. 306, Kluwer, 2001. [33] Stephen G. Simpson, Subsystems of second order arithmetic, 2nd ed., Perspectives in Logic, CUP, 2009. [34] Patrick Suppes and Rolando Chuaqui, A finitarily consistent free-variable positive fragment of Infinitesimal Analysis, Proceedings of the IXth Latin American Symposium on Mathematical Logic Notas de Logica Mathematica 38 (1993), 1-59. [35] Anne Sjerp Troelstra, Metamathematical investigation of intuitionistic arithmetic and anal- ysis, Springer Berlin, 1973. Lecture Notes in Mathematics, Vol. 344. [36] Martin V¨ath, Nonstandard analysis, Birkh¨auser Verlag, Basel, 2007. [37] F. Wattenberg, Nonstandard analysis and constructivism?, Studia Logica 47 (1988), 303–309. [38] Manfred Wolff and Peter A. Loeb (eds.), Nonstandard analysis for the working mathemati- cian, Mathematics and its Applications, vol. 510, Kluwer, 2015. Second edition.